Sunteți pe pagina 1din 30

Accepted Manuscript

The effect of anodization conditions on the morphology of porous


tungsten oxide layers formed in aqueous solution

Karolina Syrek, Leszek Zaraska, Marta Zych, Grzegorz D. Sulka

PII: S1572-6657(18)30655-6
DOI: doi:10.1016/j.jelechem.2018.09.054
Reference: JEAC 12637
To appear in: Journal of Electroanalytical Chemistry
Received date: 5 June 2018
Revised date: 29 August 2018
Accepted date: 27 September 2018

Please cite this article as: Karolina Syrek, Leszek Zaraska, Marta Zych, Grzegorz D. Sulka
, The effect of anodization conditions on the morphology of porous tungsten oxide layers
formed in aqueous solution. Jeac (2018), doi:10.1016/j.jelechem.2018.09.054

This is a PDF file of an unedited manuscript that has been accepted for publication. As
a service to our customers we are providing this early version of the manuscript. The
manuscript will undergo copyediting, typesetting, and review of the resulting proof before
it is published in its final form. Please note that during the production process errors may
be discovered which could affect the content, and all legal disclaimers that apply to the
journal pertain.
ACCEPTED MANUSCRIPT

The effect of anodization conditions on the morphology of porous tungsten oxide layers

formed in aqueous solution

Karolina Syrek*, Leszek Zaraska, Marta Zych, Grzegorz D. Sulka

Department of Physical Chemistry and Electrochemistry, Faculty of Chemistry

Jagiellonian University in Krakow, Gronostajowa 2, 30-387 Krakow, Poland

PT
RI
SC
NU
MA
T ED
EP

Keywords: tungsten oxide; anodization; nanopores; photoelectrochemical properties


C

* Corresponding author. E-mail: syrek@chemia.uj.edu.pl


AC

Department of Physical Chemistry & Electrochemistry, Faculty of Chemistry

Jagiellonian University in Krakow

Gronostajowa 2, 30-387 Krakow, Poland

Tel: +48 12 686 25 73

1
ACCEPTED MANUSCRIPT

Highlights

 Nanoporous WO 3 layers were obtained by anodization in aqueous solution.

 The effect of anodization conditions was investigated.

 An effective post-treatment procedure was also proposed.

 Conditions were optimized to obtain a highly-porous film with a maximum thickness.

 Photoelectrochemical performance of the obtained oxides was studied.

PT
Graphical abstract

RI
SC
NU
MA
ED

Abstract
T

Nanostructured WO 3 layers were obtained by one-step anodic oxidation of tungsten in


EP

an aqueous solution containing fluoride ions. A detailed investigation of the growth of anodic
C

films on metallic substrate was focused on the study of the influence of anodization
AC

conditions, such as: anodization potential, duration of the process, electrolyte composition,

applied temperature, and electrolyte agitation on morphological features of as-received

materials. Such kind of comprehensive studies has been performed for the first time. It was

found that the optimal conditions allowing synthesis of nanostructures with well-defined

pores and the thickest possible oxide layer are: anodization potential of 50 V, anodization

time of 4 h, electrolyte containing 1 M ammonium sulfate and 75 mM ammonium fluoride,

temperature of 20 °C, and agitation of electrolyte of 250 rpm. A post-treatment procedure was

2
ACCEPTED MANUSCRIPT

developed to effectively remove the precipitates formed on the surface of anodic layers during

anodization. The obtained materials were also examined as photoanodes in

photoelectrochemical (PEC) water splitting experiments. It was confirmed, that PEC

performance of nanostructured WO 3 photoanodes can be significantly improved (higher

photocurrents by 24%) after ultrasonic treatment in a 20 wt.% HF solution for 10 s and

ethanol for 5 s.

PT
.

RI
1. Introduction

SC
Tungsten oxide, WO 3 , is a well-known semiconductor characterized by a bandgap in

the range of 2.6 – 3.5 eV (depending on its structure [1]) that makes it possible to absorb even
NU
12% of the solar radiation [2]. This material also possesses very attractive optical properties,

which can be reversibly switched in the presence of electrolyte containing positively charged
MA

ions (e.g., H+, Li+ or Na+) by applying low voltages (electrochromic effect) [3]. Due to its

promising characteristics, tungsten oxide has recently found many different applications
ED

including photocatalysis [4], photoelectrochemistry [5], and gas sensing devices [6].

Moreover, nanomaterial science allows combining unique chemical, optical or electrical


T
EP

properties of a given compound with advantages of the precisely designed nanoscale

morphology, especially a high surface-to-volume ratio [7].


C

Among different synthesis strategies that can be used for the fabrication of
AC

nanostructured WO 3 [8], electrochemical oxidation of metallic tungsten is especially attractive

as it is a simple, cheap and very effective method. Up to now, this technique was successfully

applied for the formation of nanoporous or nanotubular oxide layers on either individual

metals (e.g., Al [9], Ti [10,11], W [12,13], Sn [14], Fe [15], Zr [16]) or metal alloys (TiW

[17], TiAl [18]). A great advantage of oxide layers obtained by anodization is their adhesion

to the conductive metal, as well as perpendicular orientation of the nanochannels/nanopores to

3
ACCEPTED MANUSCRIPT

the substrate that significantly facilitates an electron transfer path. This is extremely beneficial

from the point of view of their further applications (e.g., in photoelectrochemical or sensing

devices) [20]. Moreover, tailoring of the morphology of anodic films is possible by

controlling the anodization parameters, especially the applied potential, temperature, duration

of the process, as well as the composition, viscosity, and pH of the electrolyte [19].

It has been already reported, that WO 3 nanostructures can be synthesized by

PT
anodization of tungsten in various sets of conditions resulting in different oxide morphologies

RI
including nanopores [3, 12, 21–23], nanowalls/nanorods [24], nanobubbles/nanopores [25],

SC
nanoplates [26], and plate-like crystals [27]. Examples of experimental conditions used for the

synthesis of aforementioned nanostructures are collected in Table 1. It should be emphasized,


NU
that the conditions applied during anodic oxidation determine not only the type of

morphology of the resulting oxide layer, but also its geometrical parameters. For instance,
MA

WO3 layers with a pore diameter in the range of 10 – 90 nm and the oxide thickness from 0.2

μm up to 2 μm were successfully obtained in various operating conditions (see Table 1).


ED

TABLE 1

From the practical point of view (especially photocatalytic, photoelectrochemical and


T
EP

electrochemical applications), nanostructures with a highly developed surface area are the

most desirable. Therefore, establishing any correlations between anodization conditions and
C

the structural features of anodic porous WO 3 layers seems to be an extremely important issue
AC

to allow the synthesis of anodic films with desired properties like uniform pores through the

whole length of oxide, high degree of porosity, and large thickness of the resulting porous

layer. Several examples of such studies can be found in the literature. For instance, Ahmadi et

al. studied the influence of electrolyte concentration (0.01 – 0.4 M) and synthesis duration (1

– 240 min) on the morphology of anodic tungsten oxide layers formed by anodization in

oxalic acid. Surprisingly, the authors also observed the formation of a compact oxide layer

4
ACCEPTED MANUSCRIPT

with nanobubbles just underneath the typical nanoporous film [22]. A wide range of

anodization potentials was tested by Chai et al. in the same electrolyte [12]. It was found, that

both the pore diameter and thickness of anodic film increase with increasing potential applied

during anodic oxidation. As can be deduced from Table 1, very common electrolytes used for

the synthesis of porous anodic WO 3 layers are those containing F - ions [3, 23, 31]. The

influence of the fluoride ion concentration and anodization time on the morphology of anodic

PT
porous WO 3 was investigated by Lai et al. [32]. In addition, the effect of NH4 F concentration

RI
was studied by Ismail et al. [23]. It was shown that after increasing NH4 F concentration from

SC
0.5 to 0.7 g/L (14 mM to 19 mM) the thickness of resulting oxide layer remains almost the

same [23]. Some authors [24, 29, 33] conducted anodization of tungsten at elevated
NU
temperatures. Kalantar-Zadech [27] synthesized plate-like crystals with the thickness up to 60

nm and length up to 1000 nm by anodization of tungsten foil in 1.5 M HNO 3 at 50 ◦ C. Various


MA

hydrodynamic conditions during synthesis of WO 3 nanoplatelets were studied by Fernandez-

Domene et al. [33]. It was demonstrated that photoelectrochemical (PEC) response of


ED

resulting tungsten oxide can be enhanced simply by a thicker anodic film which can be

formed by controlling electrolyte stirring during the synthesis [33]. However, all
T
EP

aforementioned papers describe only the influence of some selected parameters on the

morphology of anodic WO 3 , while all other conditions are kept constant during anodic
C

oxidation. In addition, the effects of anodization temperature and electrolyte stirring on the
AC

structural features of porous WO 3 have not been reported yet. Finally, complex optimization

of all the most important operating conditions in order to obtain a highly porous anodic film

with a sufficient thickness is still lacking in the literature.

Therefore, herein we present, for the first time, a detail investigation of the effect of

such factors like: anodization potential, duration of the process, concentration of the F - ions in

the electrolyte, temperature and electrolyte agitation on the morphology of resulting

5
ACCEPTED MANUSCRIPT

nanoporous WO 3 layers. A special emphasis is focused on the optimization of all studied

operating conditions to obtain highly-porous anodic tungsten oxide with a maximal possible

thickness. In addition, we also present an easy method that can be employed to clean the

surface of anodic WO 3 from the precipitants after the synthesis process. Photoelectrochemical

properties of the received materials are also described.

2. Experimental

PT
Anodic tungsten oxide was synthesized by a single-step anodic oxidation of metallic

RI
tungsten (99.95%, 0.2 mm thick GoodFellow) carried out at the constant voltage (20 – 70 V)

in an aqueous electrolyte containing 1 M ammonium sulfate and 0 – 75 mM ammonium

SC
fluoride (every time a constant volume of electrolyte, 50 ml, was used) in the temperature
NU
range of 10 – 40 °C. The metal samples were also anodized for different durations at 50 V and

20 °C. All anodizations were performed in a typical double-wall cell (60 mm high, 40 mm in
MA

diameter) with a vertically arranged two electrodes. The W sample and the Pt plate were

served as an anode and cathode, respectively. Electrodes were kept at the constant distance of
ED

2 cm. After anodization, a post-treatment procedure was carried out by applying ultrasounds

in either ethanol or HF and HNO 3 with different concentrations for various time durations (for
T
EP

more details see Supporting Info). The morphological characterization of anodic WO 3 layers

was performed using a field emission scanning electron microscope (FE-SEM/EDS, Hitachi
C

S-4700 with a Noran System 7). Structural features of as received layers were estimated
AC

directly from SEM images using WSxM software [34]. After the post-treatment process, the

samples were annealed in air at 500 °C for 2 h with a constant heating rate of 2 °C min-1 using

a muffle furnace (model FCF 5SHM Z, Czylok) [13]. PEC measurements were performed

using a typical three-electrode cell with a quartz window. Nanostructured WO 3 photoanodes

were used as working electrodes (WEs), a platinum foil as a counter electrode (CE), and a

Luggin capillary with a saturated calomel electrode (SCE) served as a reference electrode.

6
ACCEPTED MANUSCRIPT

The generated photocurrents were measured using a photoelectric spectrometer equipped with

the 150 W xenon arc lamp and combined with a potentiostat (Instytut Fotonowy). The PEC

characterization was performed in 0.1 M KNO 3 at the potential of 1 V vs. SCE and

wavelengths ranging from 300 to 450 nm.

3. Results and discussion

3.1. Influence of anodization potential on the morphology of anodic WO3 layers

PT
As it was already mentioned, the potential applied during anodization is one of the

RI
most crucial parameters responsible for both, the kinetics of anodic layer growth and the

SC
morphology of resulting oxide. Therefore, this parameter was optimized at first. The SEM

images of the surface of anodic films obtained at different potentials are collected in Figure 1.
NU
It is clear that contrary to a quite regular and ordered morphology observed for anodic

alumina or titania, the porous layers electrochemically generated on metallic tungsten are
MA

completely irregular, independently of the potential applied during anodic oxidation. As can

be seen in Figure 1A, the anodic film with pores having a diameter of about 30 nm was
ED

obtained at the potential of 20 V (see Figure 1A). A gradual increase in the average pore size

with increasing anodization potential was observed. In consequence, well defined and
T
EP

relatively uniform pores having about 90 nm in diameter were created when the oxide layer

was synthesized at 50 V (Figure 1D). A similar dependence between the pore size and applied
C

potential, typical for other porous anodic metal oxides [19], were also observed by Chai et al.
AC

for anodizations of tungsten carried out in the oxalic acid electrolyte [12]. However, as it is

clearly visible in Figure 1E and F, a further increase in applied voltage to 60 V or 70 V results

in the formation of anodic films with clogged pore mouths. This can be attributed to the fact

that WO3 dissolution also occurs during anodization according to the following reactions:

𝑊𝑂3 + 2𝐻 + → 𝑊𝑂22+ + 𝐻2 𝑂 (1)

𝑊𝑂3 + 6𝐻 + + 8𝐹 − → [𝑊𝐹8 ]2− + 3𝐻2 𝑂 (2)

7
ACCEPTED MANUSCRIPT

After reaching supersaturation conditions, a precipitate of hydrated tungsten oxide (WO 3 ·H2 O

or WO3 ·2H2 O) may be deposited on the surface of anodic film [33,35]. Therefore, from the

point of view of the surface topography, 50 V was found to be the highest possible potential

allowing the formation of well-defined nanoporous tungsten oxide layers in the studied

electrolyte.

As stated previously, the fabrication of the nanostructured WO 3 layer with a sufficient

PT
thickness ensuring the optimal photoelectrochemical performance of the photoanode is still a

RI
challenge. Therefore, the effect of anodization potential on the thickness of the resulting oxide

SC
film was also investigated in order to find the optimal value of this parameter for the

formation of the thickest possible anodic film.


NU
FIGURE 1

The cross-sectional SEM images of the anodic film grown at the potential of 50 V are shown
MA

in Figures 2A and B. As can be seen, the morphology of the as formed porous tungsten oxide

layer with characteristic bubbles formed at the metal/oxide interface (Figure 2B), is
ED

comparable to those previously obtained in similar [36], as well as completely different

electrolytes (for details see Table 1) [25, 36].


T
EP

FIGURE 2

Thicknesses of anodic films synthesized at all studied potentials were determined from the
C

cross-sectional SEM images, and the results are collected in Figure 2D (black curve). It is
AC

clear that the increasing anodization potential from 20 V to 60 V results in the formation of

gradually thicker oxide layers, and this dependence is almost linear. On the other hand, a

further rising of the potential to 70 V leads to the generation of the significantly thinner

anodic film, what can be related to the worse current efficiency of the oxide formation itself,

as well as enhanced oxide dissolution under the electric field resulting in the formation of

interconnected pores. Moreover, the anodic layer can be partially damaged by vigorous

8
ACCEPTED MANUSCRIPT

oxygen evolution. An oxide growth ratio, defined as an average thickness of the oxide layer

divided by the density of charge passed during anodization [37], was calculated for all

samples (blue curve in Figure 2D). The charge densities were obtained by integrating the

current-density vs. time curves. Indeed, the maximum value of the oxide growth ratio was

observed for the sample anodized at 60 V suggesting the highest efficiency of the process

carried out at these conditions. Applying the higher anodization potential results in much

PT
effective oxide dissolution caused by a high electric field [38] as well as vigorous oxygen

RI
evolution which is accompanied by a relatively high electronic current. In consequence, high

SC
and fluctuating current densities are observed (see Figure S1 in the Supporting Information).

It should be also mentioned that the current density vs. time curves recorded during
NU
anodizations carried out at the potentials lower than 60 V (Figure 2C), exhibit a typical shape

observed during a potentiostatic growth of porous oxide films on different metals. Taking
MA

above into account, the potential of 50 V, ensuring both a well-defined porous surface

morphology together with the maximum possible thickness of the anodic film, was chosen as
ED

optimal for further investigations.


T

3.2. The effect of anodization duration


EP

The second parameter that has been taken into consideration was anodization time. A

series of anodizations was carried out at the optimal potential of 50 V for various durations
C

ranging from 15 min to 360 min. The SEM images of resulting anodic layers are collected in
AC

Figure 3 and Figure S2 (see Supporting Information). As can be seen in Figure 3A, even after

180 min of anodic oxidation, the resulting oxide layer still does not exhibit a well-defined

porous morphology. Some pores are still clogged on the surface. This can suggest that the

time was too short to achieve effective etching of the compact oxide layer created at the early

stages of the anodization. When the anodization was prolonged to 240 min, distinct pores

appear on the surface of anodic film (Figure 3B). However, a further increase of the

9
ACCEPTED MANUSCRIPT

anodization time to 360 min results again in almost compact morphology due to the chemical

dissolution of the anodic oxide by F - ions. The oxide layer thicknesses and oxide growth

ratios were calculated for all studied samples, and the obtained results are shown in Figure 3D

(black and blue curves, respectively). It is clear, that the most effective oxide layer thickening

occurs at the initial stage of anodization (see the highest growth ratio for the sample anodized

for 15 min). On the other hand, the longer the process, the chemical etching of the formed

PT
structure becomes a predominant factor preventing the formation of significantly thicker

RI
layers [39]. Since an extension of time of anodization from 240 to 360 min results in the

SC
formation of a slightly thicker layer (Figure 3D) and, simultaneously, in the loss of well-

defined porous morphology (Figure 3C), 240 min was chosen as an optimal process duration
NU
for further investigations.

FIGURE 3
MA

3.3. The effect of anodization temperature


ED

Anodization temperature is another very important parameter that can significantly

affect the morphology of anodic tungsten oxide layers [27]. Therefore, metallic tungsten was
T
EP

also anodized for 240 min at the optimal potential of 50 V using four different electrolyte

temperatures ranging from 10 to 40 °C. The SEM images of resulting oxide layers are
C

collected in Figure 4 together with current density vs. time curves (Figure 4E), thicknesses of
AC

anodic films, and values of oxide growth ratio determined for all studied samples (Figure 4F).

FIGURE 4

The thickest oxide layer with a thickness of about 800 nm was obtained when anodic

oxidation was carried out at the temperature of 10 °C. However, as shown in Figure 4A, the

resulting oxide layer was rather compact, with some irregular cavities and holes on the

surface. As can be seen in Figure 4E, increasing anodization temperature results in a

10
ACCEPTED MANUSCRIPT

significant change in current density vs. time curves. It is clear that the higher temperature,

the initiation of pore formation occurs faster resulting in an earlier increase in current density

after reaching the local minimum (see Figure 4E). The higher current densities observed

during the steady-state oxide growth can suggest the faster oxide formation at elevated

temperatures (see Figure S3 in the Supporting Information). A similar trend was observed in

our group for the anodic formation of nanostructured oxides on other metals such as

PT
aluminum [40] and titanium [41]. However, in the present case, increasing anodization

RI
temperature leads to the formation of much thinner anodic films (see Figure 4F). This

SC
phenomenon can be attributed to the fact that the rate of chemical dissolution of oxide is

increased more by raising the temperature than by the rate of oxide growth. In consequence,
NU
the oxide dissolution is predominant process at higher temperatures that results in the

formation of thinner layers. On the other hand, more effective etching facilitates the pore
MA

formation within the anodic film as can be seen in Figure 4B. However, when anodization is

carried out at too high temperatures, the anodic layers with less regular and sometimes
ED

clogged pores are formed due to the deposition of hydrated tungsten oxide (see above).

Taking above into account, 20 °C was chosen as the optimal temperature for further studies.
T
EP

3.4. The effect of fluoride ion concentration

As the dissolution of anodic tungsten oxide occurs according to the equation (2), the
C

concentration of F- ions in the electrolyte is also an extremely important parameter affecting


AC

the morphology and thickness of the resulting film. Therefore, we decided to check if the use

of electrolyte containing less or even no fluoride ions can allow the formation of thicker

porous layers. Despite the fact that some porous anodic film can be obtained on the tungsten

surface even in the solution without fluoride ions, this electrolyte is not promising, as the

formed oxide layer is extremely thin (~240 nm) and mostly compact with only thin porous

outer layer (SEM pictures are shown in Figure S4 in the Supporting Information). It is clearly

11
ACCEPTED MANUSCRIPT

visible that increasing the concentration of NH4 F results in higher current densities observed

during anodization (Figure 5D) due to the better conductivity of electrolyte. Moreover, the

higher concentration of fluoride ions facilitates the pore formation, as indicated by the earlier

increase (after the local minimum) in the current density. Surprisingly, on the contrary to

results obtained by Ismail et al. [23], the thickest anodic layer was obtained when the

anodization was carried out in the electrolyte with the highest content of F- ions (see Figure

PT
5E). However, the more F- ions present the greater rate of oxide dissolution that it is

RI
confirmed by much lower oxide growth ratios in the concentrated electrolytes (see Figure 5E).

It should be also mentioned that if the concentration of F - is too low, porous oxide films with

SC
not fully open pores are formed, as can be seen in Figures 5A and C. Therefore, 75 mM was
NU
chosen as the optimal concentration of NH4 F allowing the fabrication of porous anodic

tungsten oxide layers with well-defined pores and the greatest possible thickness.
MA

FIGURE 5

3.5. Influence of electrolyte agitation


ED

Several studies showed a big impact of electrolyte agitation on the thickness and

morphology of anodic oxides layers, e.g., TiO2 [42,43]. Therefore, we also decided to
T
EP

optimize the speed of electrolyte stirring to achieve a highly porous and the thickest possible

tungsten oxide film.


C

FIGURE 6
AC

As can be seen in Figure 6A, when anodic oxidation was carried out without electrolyte

agitation, pores were not fully developed on the surface. A large area of the compact product,

being either not fully dissolved passive oxide formed during the initial stages of anodization

or hydrated oxide secondary deposited from the electrolyte, can be easily recognized. On the

other hand, a gradual decrease in the oxide layer thickness as well as in the oxide growth ratio

was observed for increasing the speed of electrolyte stirring (see Figure 6D). Therefore, it can

12
ACCEPTED MANUSCRIPT

be stated, that 250 rpm is the optimal agitation rate for the fabrication of porous anodic

tungsten oxide layers in the studied electrolyte.

3.6. Post-treatment procedure

In order to clean the anodic WO 3 surface from post-anodization precipitates different

approaches were examined. Firstly, just after anodization anodic tungsten oxide was

immersed in a mixture of HF and water of different ratios. This approach was not successful,

PT
as no considerably differences were observed between the samples before and after post-

RI
treatment (for details see section 5.1 in the Supporting Information). Secondly, as-received

SC
WO3 samples were immersed in ethanol and exposed to ultrasounds for different times up to

10 min. A similar strategy based on ultrasonic treatment has been already used to clean the
NU
surface of TiO 2 nanotubes after their synthesis [44,45]. In case of anodic WO 3 layers some

improvement in a surface quality was observed, however, the surface was still not entirely
MA

free from the precipitates (for details see section 5.2 in the Supporting Information). Finally,

the samples were treated using a two step method based on the sequential immersion of
ED

samples in HF and/or HNO 3 with different concentrations, followed by ethanol in the


T

presence of ultrasounds (for details see section 5.3 in the Supporting Information). This
EP

approach allowed finding the appropriate conditions for removal of all precipitates from WO 3

surface. The best results were obtained by sonication in a 20% HF solution for 10 s and
C

ethanol for 5 s. The SEM images of the surface of anodic film before and after the post-
AC

treatment procedure are shown in Figure 7A and B, respectively.

3.7. Photoelectrochemical properties of anodic WO3

As it was already mentioned, among many potential applications of anodic WO 3 ,

photocatalysis (PC) and photoelectrochemical (PEC) water splitting seem to be especially

perspective. Detail studies on PEC performance of anodic WO3 photoanodes have been

already performed in our group [13]. It is widely recognized that PEC properties are strongly

13
ACCEPTED MANUSCRIPT

related with the morphology of anodic layers. There are several studies which prove that a top

nanotube reflection is a very important factor affecting photochemical and

photoelectrochemical properties [43]. In order to check the effect of post-treatment procedure

on PEC performance of porous anodic WO 3 , just after the anodization one sample was cut in

half, and one piece was left as received and the second was subjected to the post treatment

procedure. Prior to photoelectrochemical measurements both specimens were annealed at 500

PT
°C for 2 h. The PEC properties of both samples were studied at the potential of 1 V vs. SCE

RI
by a sequential irradiation of light from the wavelength range of 300 – 450 nm with the step

SC
of 10 nm. The obtained chronoamperograms are presented in Figure 7C. It is clearly visible

that PEC response of the WO3 photoanode subjected to the post-treatment procedure was
NU
higher of about 22% compared to the as-received sample. For 350 nm and 1 V vs. SCE, the

observed average enhancement in photocurrent was 24 ± 3 % (for details see Figure S8 in the
MA

Supporting Information).

FIGURE 7
ED

Conclusions

In summary, nanoporous tungsten oxide layers with a thickness of several hundred of


T
EP

nanometers were successfully obtained by one-step anodization in the electrolyte containing

ammonium sulfate and ammonium fluoride. The main goal of this study was to find an
C

optimal procedure for obtaining WO3 layers with a desired morphology – well-defined pores
AC

and the thickest possible oxide layer. As expected, we demonstrated a significant influence of

several anodization parameters on the resulting morphology of anodic films. The detailed

analysis of the established correlations between the anodization parameters and oxide

morphology allowed choosing the optimal conditions for the synthesis of anodic WO 3 in this

electrolyte. The best conditions can be summarized as follows: the applied potential of 50 V,

anodization time of 4 h, electrolyte containing 1 M ammonium sulfate and 75 mM ammonium

14
ACCEPTED MANUSCRIPT

fluoride, anodization temperature of 20 °C, and the rate of electrolyte agitation of 250 rpm.

Finally, the post-treatment procedure based on ultrasonication in a HF solution for 10 s and

then in ethanol was proposed as an effective strategy for the removal of precipitates formed

during the anodization from the porous surface of anodic oxide. It was confirmed that such a

cleaning process can enhance the photoelectrochemical properties of obtained materials. A

24% increase in generated photocurrents was indicated.

PT
RI
Acknowledgements

SC
The research was partially supported by the National Science Centre Poland (Project

No. 2016/23/B/ST5/00790). Karolina Syrek acknowledges support from National Science


NU
Centre (Poland) through the doctoral scholarship program ETIUDA4, grant number

2016/20/T/ST5/00255. The SEM imaging was performed in the Laboratory of Field Emission
MA

Scanning Electron Microscopy and Microanalysis at the Institute of Geological Sciences,

Jagiellonian University, Poland.


T ED

References
EP

[1] O. Carp, C.L. Huisman, A. Reller, Photoinduced reactivity of titanium dioxide, Prog.

Solid. State Chem. 32 (2004) 33–177.


C

[2] S.S. Kalanur, Y.J. Hwang, S.Y. Chae, O.S. Joo, Facile growth of aligned WO 3 nanorods
AC

on FTO substrate for enhanced photoanodic water oxidation activity, J. Mater. Chem. A 10

(2013) 3479–3488.

[3] J.Z. Ou, S. Balendhran, M. R. Field, D. G. McCulloch, A.S. Zoolfakar, R. A. Rani, S.

Zhuiykov, A. P. O’Mullane, K. Kalantar-Zadeh, The anodized crystalline WO 3 nanoporous

network with enhanced electrochromic properties, Nanoscale 4 (2012) 5980–5988.

15
ACCEPTED MANUSCRIPT

[4] A. Watcharenwong, W. Chanmanee, N.R. de Tacconi, C.R. Chenthamarakshan,

P.Kajitvichyanukul, K. Rajeshwar, Anodic growth of nanoporous WO 3 films: Morphology,

photoelectrochemical response and photocatalytic activity for methylene blue and hexavalent

chrome conversion, J. Electroanal. Chem. 612 (2008) 112–120.

[5] A. Tacca, L. Meda, G. Marra, A. Savoini, S. Caramori, V. Cristino, C.A. Bignozzi, V.G.

Pedro, P.P. Boix, S. Gimenez, J. Bisquert, Photoanodes based on nanostructured WO 3 for

PT
water splitting, ChemPhysChem 13 (2012), 3025–3034.

RI
[6] W. Zeng, Y. Li, B. Miao, K. Pan, Hydrothermal synthesis and gas sensing properties of

WO3 ∙H2 O with different morphologies, Physica E 56 (2014) 183–188.

SC
[7] M. Paulose, K. Shankar, S. Yoriya, H.E. Prakasam, O.K. Varghese, G.K. Mor, T.A.
NU
Latempa, A. Fitzgerald, C.A. Grimes, (2006) Anodic growth of highly ordered TiO 2 nanotube

arrays to 134 μm in length, J. Phys. Chem. B 110 (2006) 16179–16184.


MA

[8] Z. Liu, M. Miyauchi, T. Yamazaki, Y. Shen, Facile synthesis and NO 2 gas sensing of

tungsten oxide nanorods assembled microspheres, Sensors Actuat. B 140 (2009) 514–519.
ED

[9] L. Zaraska, A. Brudzisz, E. Wierzbicka, G.D. Sulka, The effect of electrolyte change on

the morphology and degree of nanopore order of porous alumina formed by two-step
T
EP

anodization, Electrochim. Acta 198 (2016) 259–267.

[10] G.D. Sulka, J. Kapusta-Kołodziej, A. Brzózka, M. Jaskuła, Fabrication of nanoporous


C

TiO2 by electrochemical anodization, Electrochim. Acta 55 (2010) 4359–4367.


AC

[11] J.M. Macak, H. Hildebrand, U. Marten-Jahns, P. Schmuki, Mechanistic aspects and

growth of large diameter self-organized TiO 2 nanotubes, J. Electroanal. Chem. 621 (2008)

254–266.

[12] Y. Chai, C.W. Tam, K.P. Beh, F.K. Yam, Z. Hassan, Porous WO 3 formed by anodization

in oxalic acid, J. Porous Mater. 20 (2013) 997–1002.

16
ACCEPTED MANUSCRIPT

[13] K. Syrek, M. Zych, L. Zaraska, G.D. Sulka, Influence of annealing conditions on anodic

tungsten oxide layers and their photoelectrochemical activity, Electrochim. Acta 231 (2017)

61–68.

[14] L. Zaraska, K. Gawlak, M. Gurgul, D.K. Chlebda, R.P. Socha, G.D. Sulka, Controlled

synthesis of nanoporous tin oxide layers with various pore diameters and their

photoelectrochemical properties, Electrochim. Acta 254 (2017) 238–245.

PT
[15] A. Pawlik, K. Hnida, R.P. Socha, E. Wiercigroch, K. Małek, G.D. Sulka, Effects of

RI
anodizing conditions and annealing temperature on the morphology and crystalline structure

SC
of anodic oxide layers grown on iron, Appl. Surf. Sci. 426 (2017) 1084–1093.

[16] E. Wierzbicka, K. Syrek, G.D. Sulka, M. Pisarek, M. Janik-Czachor, The effect of foil
NU
purity on morphology of anodized nanoporous ZrO 2 , Appl. Surf. Sci. 388 (2016) 799 – 804.

[17] J. Zhao, X. Wang, Y. Kang, X. Xu, Y. Li, Photoelectrochemical activities of W-doped


MA

titania nanotube arrays fabricated by anodization, IEEE Photonic. Tech. L. 20 (2008) 1213–

1215.
ED

[18] S. Berger, H. Tsuchiya, P. Schmuki, Transition from nanopores to nanotubes: self-

ordered anodic oxide structures on titanium-aluminides, Chem. Mater. 20 (2008) 3245–3247.


T
EP

[19] A. Ghicov, P. Schmuki, Self-ordering electrochemistry: a review on growth and

functionality of TiO 2 nanotubes and other self-aligned MO(x) structures, Chem. Commun. 20
C

(2009) 2791–2808.
AC

[20] P. Xiao, D. Liu, B.B. Garcia, S. Sepehrib, Y. Zhang, G. Cao, Electrochemical and

photoelectrical properties of titania nanotube arrays annealed in different gases, Sensors

Actuat. B 134 (2008) 367–372.

[21] M. Altomare, O. Pfoch, A. Tighineanu, R. Kirchgeorg, K. Lee, E. Selli, P. Schmuki,

Molten o-H3 PO4 : a new electrolyte for the anodic synthesis of self-organized oxide structures-

WO3 nanochannel layers and others, J. Am. Chem. Soc. 137 (2015) 5646–5649.

17
ACCEPTED MANUSCRIPT

[22] E. Ahmadi, C.Y. Ng, K.A. Razak, Z. Lockman, Preparation of anodic nanoporous WO 3

film using oxalic acid as electrolyte, J. Alloy. Compd. 704 (2017) 518–527.

[23] S. Ismail, C.Y. Ng, E. Ahmadi, K.A. Razak, Z. Lockman, Segmented nanoporous WO 3

prepared via anodization and their photocatalytic properties, J. Mater. Res. 31 (2016) 712–

728.

[24] J. Zhang, I. Salles, S. Pering, P.J. Cameron, D. Mattia, S. Eslava, Nanostructured WO 3

PT
photoanodes for efficient water splitting via anodisation in citric acid, RSC Adv. 7 (2017)

RI
5221–35227.

SC
[25] W-H. Chen, M-Y. Lai, K-T. Tsai, C-Y. Liu, Y-L. Wang, Spontaneous formation of

ordered nanobubbles in anodic tungsten oxide during anodization, J. Phys. Chem. C 115
NU
(2011) 18406–18411.

[26] Chin wei Lai, Tungsten Oxide Nanoplate Film as a Highly Active Solar-Driven
MA

Photocatalyst, Curr.Nanosci. 2015, 11(999) 1-1.

[27] K. Kalantar-zadeh, A.Z. Sadek, H. Zheng, V. Bansal, S. K. Bhargava, W. Wlodarski, J.


ED

Zhu, L. Yu, Z. Hu, Nanostructured WO 3 films using high temperature anodization, Sensors

Actuat. B 142 (2009) 230–235.


T
EP

[28] K.K. Upadhyay, M. Altomare, S. Eugénio, P. Schmuki, T.M. Silva, M.F. Montemor, On

the supercapacitive behaviour of anodic porous WO 3 -based negative electrodes, Electrochim.


C

Acta 232 (2017) 192–201.


AC

[29] R.M. Fernández-Domene, R. Sánchez-Tovar, B. Lucas-Granados, G. Roselló-Márquez,

J. García-Antón, A simple method to fabricate high-performance nanostructured WO 3

photocatalysts with adjusted morphology in the presence of complexing agents, Mater. Des.

116 (2017) 160–170.

18
ACCEPTED MANUSCRIPT

[30] C-K. Wang, C-K. Lin, C-L. Wu, S-C. Wang, J-L. Huang, Synthesis and characterization

of electrochromic plate-like tungsten oxide films by acidic treatment of electrochemical

anodized tungsten, Electrochim. Acta 112 (2013) 24– 31.

[31] Q. Zheng, H-J. Lee, J. Lee, W. Choi, N-B. Park, C. Lee, Electrochromic titania nanotube

arrays for the enhanced photocatalytic degradation of phenol and pharmaceutical compounds,

Chem. Eng. J. 249 (2014) 285–292.

PT
[32] Chin Wei Lai, Photocatalysis and Photoelectrochemical Properties of Tungsten Trioxide

RI
Nanostructured Films, Sci. World J. (2014) 843587-843594.

SC
[33] R.M. Fernández-Domene, R. Sánchez-Tovar, E. Segura-Sanchís, J. García-Antón, Novel

tree-like WO 3 nanoplatelets with very high surface area synthesized by anodization under
NU
controlled hydrodynamic conditions, Chem. Eng. J. 286 (2016) 59–67.

[34] I. Horcas, R. Fernández, J.M. Gómez-Rodríguez, J. Colchero, J. Gómez-Herrero, A.M.


MA

Baro, WSxM: a software for scanning probe microscopy and a tool for nanotechnology, Rev.

Sci. Instrum. 78 (2007) 1-8.


ED

[35] R.M. Fernández-Domene, R. Sánchez-Tovar, B. Lucas-Granados, J. García-Antón,

Improvement in photocatalytic activity of stable WO 3 nanoplatelet globular clusters arranged


T
EP

in a tree-like fashion: Influence of rotation velocity during anodization, Appl. Catal. B-

Environ. 189 (2016) 266–282.


C

[36] C.Y. Ng, K.A. Razak, Z. Lockman, Effect of annealing temperature on anodized
AC

nanoporous WO 3 , J. Porous Mater. 22 (2015) 537–544.

[37] L. Zaraska, K. Gawlak, M. Gurgul, M. Dziurka, M. Nowak, D. Gilek, G.D. Sulka,

Influence of anodizing conditions on generation of internal cracks in anodic porous tin oxide

films grown in NaOH electrolyte, Appl. Surf. Sci. 439 (2018) 672–680.

19
ACCEPTED MANUSCRIPT

[38] A. Ghicov, P. Schmuki, Self-ordering electrochemistry: a review on growth and

functionality of TiO 2 nanotubes and other self-aligned MO x structures, Chem. Commun.

(2009) 2791–2808.

[39] M. Yang, N. K. Shrestha, P. Schmuki, Thick porous tungsten trioxide films by

anodization of tungsten in fluoride containing phosphoric acid electrolyte, Electrochem.

Commun. 11 (2009) 1908–1911.

PT
[40] L. Zaraska, G.D. Sulka, M. Jaskuła, Anodic alumina membranes with defined pore

RI
diameters and thicknesses obtained by adjusting the anodizing duration and pore

opening/widening time, J. Solid State Electrochem. 15 (2011) 2427 – 2436.

SC
[41] J. Kapusta-Kołodziej, O. Tynkevych, A. Pawlik, M. Jarosz, J. Mech, G.D. Sulka,
NU
Electrochemical growth of porous titanium dioxide in a glycerol-based electrolyte at different

temperatures, Electrochim. Acta 144 (2014) 127–135.


MA

[42] K. Syrek, J. Kapusta-Kołodziej, M. Jarosz, G.D. Sulka, Effect of electrolyte agitation on

anodic titanium dioxide (ATO) growth and its photoelectrochemical properties, Electrochim.
ED

Acta 180 (2015) 801 – 810.

[43] R. Sánchez-Tovar, I. Paramasivam, K. Lee, P. Schmuki, Influence of hydrodynamic


T
EP

conditions on growth and geometry of anodic TiO 2 nanotubes and their use towards optimized

DSSCs, J. Mater. Chem. 25 (2012) 12792-12795.


C

[44] H. Xu, Q. Zhang, C. Zheng, W. Yana, W. Chu, Application of ultrasonic wave to clean
AC

the surface of the TiO 2 nanotubes prepared by the electrochemical anodization, Appl. Surf.

Sci. 257 (2011) 8478– 8480.

[45] C.W. Lai, S. Sreekantan, Photoelectrochemical performance of smooth TiO 2 nanotube

arrays: Effect of anodization temperature and clearing methods, Int. J. Photoenergy (2012)

356943- 356954.

20
ACCEPTED MANUSCRIPT

Figure captions

Figure 1. Top-view SEM images of anodic WO 3 layers formed in 1 M (NH4 )2 SO4 and 75 mM

NH4 F at the potential of 20 V (A), 30 V (B), 40 V (C), 50 V (D), 60 V (E) and 70 V (F). All

anodizations were carried out at 20 °C for 4 h.

Figure 2. Cross-sectional SEM images of the nanostructured tungsten oxide layer formed at

the potential of 50 V (A – lower magnification, B – higher magnification). Current density vs.

PT
time curves (C) recorded at various anodization potentials. The oxide thickness and oxide

RI
growth ratio as functions of the applied potential (D).

SC
Figure 3. Top-view SEM images of anodic WO 3 layers formed in 1 M (NH4 )2 SO4 and 75 mM

NH4 F at 50 V for 180 min (A), 240 min (B), and 360 min (C). The oxide growth ratio and
NU
resulting oxide thickness for various anodization durations (D).

Figure 4. Top-view SEM images of WO 3 layers synthesized in 1 M (NH4 )2 SO 4 and 75 mM


MA

NH4 F at 50 V for 240 min at 10 °C (A), 20 °C (B), 30 °C (C), and 40 °C (D). Current density

vs. time curves recorded during anodizations at different temperatures (E). The oxide growth
ED

ratio and resulting oxide thickness as functions of anodization temperature (F).

Figure 5. SEM images of nanostructured WO 3 layers obtained by anodization of metallic


T
EP

tungsten at 50 V and 20 °C in the electrolyte containing 33 mM (A), 54 mM (B), and 75 mM

(C) of NH4 F. The corresponding current density-time curves recorded during synthesis (D).
C

Oxide growth ratios and resulting oxide thicknesses for all studied F - ion concentrations (E).
AC

Figure 6. SEM images of WO 3 layers synthesized at 0 rpm (A), 250 rpm (B), and 500 rpm

(C). Oxide growth ratios and resulting oxide thicknesses for all studied agitation speeds (D).

Figure 7. Comparison of the top morphology of anodic WO 3 before (A) and after the post-

treatment process performed in ultrasounds (B), together with the corresponding PEC

responses (C).

21
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
T ED
EP
C

Fig. 1
AC

22
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA

Fig. 2
T ED
EP
C
AC

23
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA

Fig. 3
T ED
EP
C
AC

24
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
T ED
EP
C

Fig. 4
AC

25
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
T ED
EP
C
AC

Fig. 5

26
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA

Fig. 6
T ED
EP
C
AC

27
ACCEPTED MANUSCRIPT

PT
RI
SC
NU
MA
T ED
EP

Fig. 7
C
AC

28
ACCEPTED MANUSCRIPT

Table 1. Structural features of anodic tungsten oxide. Where: RT – room temperature.


Structural parameters
Electrolyte composition Applied voltage Time of synthesis Temperature
Pore size Wall thickness Oxide thickness Ref.
(WO3 morphology) (V) (min) (°C)
(nm) (nm) (µm)
pure molten ortho-phosphoric 1 – 20 60 100 10 - 0.19 – 1.3 [21]
acid
(nanopores)
5 60 – 240 100 10
P T - 0.2 – 0.8 [29]

0.1 M citric acid in H2 O


*constant
R I total: 3.7 ± 1.5
(0.5
(nanowalls/nanorods)
current:
0.05 – 0.1 A
30 0

S C 100 – 200 nanostructured


valleys)
[24]
0.1 M citric acid in solution of 80
vol% NMF and 20 vol% H2 O 40 360 40
N U 50 - 5.8 ± 1.7
(randomly oriented nanowires)
1.5 M H2 SO4 and NaF in the
range 0 – 0.25 M
M A - - ~ 1.0 – 16

D
(nanoplates)
20 240 50 [29]
1.5 M H2 SO4 and H2 O2 in the
range 0 – 0.2 M
(nanosheets)
T E - - ~ 1.0 – 2.5

0.3 M oxalic acid


(nanopores)
20 – 100

E P 30 RT 50 – 90 - 0.1 – 0.4 [12]

C
1 M Na2 SO4 and NH4 F (0.07, 0.1,
0.3, 0.5, and 0.7 g/L) 40 60 RT 50 – 80 - ~ 0.4 [23]
(nanopores)
50 ml ethylene glycol with 0.5 g
NH4 F and 2 vol.% water A C 10 10 – 90 RT 10 – 30 few nm 0.2 – 2.0 [3]
(nanopores)
0.2 wt.% NH4 F
60 18 RT 80 – 100 0.04 – 0.08 [30]
(nanoplates)
0.25 wt.% HF
50 40 RT 50 – 70 15 - [31]
(nanopores)

29

S-ar putea să vă placă și